Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Spatial resolution in optical coherence elastography of bounded media

Open Access Open Access

Abstract

Dynamic optical coherence elastography (OCE) tracks mechanical wave propagation in the subsurface region of tissue to image its shear modulus. For bulk shear waves, the lateral resolution of the reconstructed modulus map (i.e., elastographic resolution) can approach that of optical coherence tomography (OCT), typically a few tens of microns. Here we perform comprehensive numerical simulations and acoustic micro-tapping OCE experiments to show that for the typical situation of guided wave propagation in bounded media, such as cornea, the elastographic resolution cannot reach the OCT resolution and is mainly defined by the thickness of the bounded tissue layer. We considered the excitation of both broadband and quasi-harmonic guided waves in a bounded, isotropic medium. Leveraging the properties of broadband pulses, a robust method for modulus reconstruction with minimum artifacts at interfaces is demonstrated. In contrast, tissue bounding creates large instabilities in the phase of harmonic waves, leading to serious artifacts in modulus reconstructions.

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

The speed of a propagating mechanical wave in a medium is fully defined by its complex elastic modulus and the boundary conditions. Often, boundary conditions are ignored (i.e., bulk plane-wave propagation in an infinite medium is assumed) in dynamic elastography such that the mechanical modulus at every sample point along a propagation path can be fully defined by the measured local wave speed. If the assumption of plane wave propagation in a bulk medium is accurate, spatially resolved mechanical modulus reconstruction is a simple one-to-one mapping from the locally measured wave speed. Indeed, many soft biological tissues are nearly incompressible and are characterized by a single elastic modulus, which can be computed from the measured shear wave speed through a simple expression [13].

For the ideal case where plane shear waves are used and interface waves can be ignored, the spatial resolution of the reconstructed modulus is limited by the capabilities of the imaging system. For example, spatial resolution in conventional ultrasound elastography (USE) is primarily determined by the scanning beam width (on the order of a few hundreds of micrometers) [4], while optical coherence tomography (OCT) has a lateral resolution determined by the optical focusing power of the scanning lens and can be on the order of microns [5]. OCT-based elastography (OCE) holds the promise for high-resolution modulus reconstruction due to its high-resolution imaging, non-contact nature, and ability to detect nanometer scale vibrations [1,611]. Compared to USE, OCE has the potential to improve elastographic resolution by an order of magnitude. Measured resolution in wave-based OCE has never approached the ultimate resolution of the OCT system, however, even though OCE resolution is often assumed to equal that of the OCT imaging system.

In this work, we define elastographic resolution as the minimum inclusion size for which the mechanical modulus can be accurately reconstructed with little or no a-prior knowledge of material stiffness. OCE usually images shallow subsurface tissue regions and, therefore, probes surface propagating mechanical waves. As shown in [1], the mechanical resolution in dynamic OCE utilizing surface (or Rayleigh) mechanical waves is generally limited by the characteristics of the mechanical wave (frequency, bandwidth) and not by the capabilities of the OCT system (assuming sufficient temporal and spatial sampling for OCT reconstruction of light scattering and wavefront propagation). In heterogeneous materials, Rayleigh waves mix with different kinds of waves generated at boundaries, for example at the edges of an inclusion, that complicate wave analysis. Mode conversions and guided propagation can lead to serious errors in the reconstructed elastic modulus [1,2]. As shown previously, the lateral resolution in elastic modulus maps reconstructed from surface wave propagation captured with dynamic OCE is about half of the mechanical wavelength, or on the order of the mechanical pulse width (for broadband signals) [1].

Because OCT is generally restricted to surface measurements (on the order of 2-3 mm in depth), and many tissues have a layered structure, wave-based OCE methods do not exclusively use Rayleigh waves, i.e. waves propagating over the surface of a bulk material. Many tissues and organs (skin, cornea, gastro-intestinal channels, etc.) are bounded, leading to guided mechanical waves [35]. Guided waves are aggregate waves formed by reflections and reverberations of bulk mechanical waves inside a bounded layer. As shown previously, guided elastic waves are dispersive and require an appropriate mechanical model to reconstruct elastic moduli from them [2,6].

In this study, we conducted numerical simulations in OnScale (Redwood City, California) and performed acoustic micro-tapping (A$\mu $T) OCE measurements to explore the spatial resolution of reconstructed modulus maps for a layered, isotropic medium where the thickness is on the order of the mechanical wavelength (i.e., guided wave propagation). The lateral resolution in quantitative modulus images is limited by the wavelength of the elastic wave, similar to the Rayleigh wave case, but additionally depends on the layer thickness.

2. Materials and methods

2.1. Numerical simulations

A finite element (FEM) simulation was designed to simulate ideal, noise-free guided wave behavior in a two-part isotropic medium (Fig. 1). The simulation geometry is shown in Fig. 1(a). The lengths of the first and second parts are 10 mm and 20 mm, respectively. The shear modulus of part I was held constant at $G = {G_1} = 30$ kPa, and the shear modulus of part II was set to different values of $G = {G_2} = [{30,\,33,\,45,\,60,\,120} ]$ kPa for different simulations. The material density ($\rho $ = 1000 kg/m3) was uniform throughout both parts. The longitudinal wave speed was uniform (${c_L}$ = 1540 m/s) throughout the entire domain (including water) except for air. Four different thicknesses $h = [{0.125,\; 0.25,\; 0.5,\; 1} ]$ mm were investigated.

 figure: Fig. 1.

Fig. 1. Numerical simulation in OnScale of guided wave propagation in a two-part nearly incompressible linear isotropic material of uniform thickness ($h = 0.5$ mm in this figure) bounded below by water and above by air (mimicking cornea boundary conditions). Shear moduli in parts I and II used in this figure are equal to $G = {G_1} = 30\; $kPa and $G = {G_2} = 60\; $kPa respectively; medium thickness is uniform over the medium and equal to $h = 0.5$ mm. The interface between the two parts is at $x = 10$ mm. Broadband pulses of guided waves were launched with acoustic micro-tapping [10] from air. The push profile was considered Gaussian in space and super-Gaussian in time, as described in Eq. (1). (a) Medium geometry used in the OnScale simulations. A 1-mm layer of water bounds the media on the right-hand side to better match absorbing boundary conditions. (b) Space-time ($XT$) plot of guided wave propagation at the medium surface. (e) Wavenumber-frequency ($kf$) spectrum obtained from $XT$ plot (b) using an 8 mm window in part II with solutions for ${A_0}$ and ${S_0}$ modes best fitting the spectrum. (h) Goodness of fit (GOF) of the simulated spectrum as a function of ${G_1}$ describing the method of shear modulus reconstruction from the $kf$ spectrum. (c), (f), (i) are $XT$ plots with a 4 mm window, $kf$ spectrum and GOF respectively computed for this window. (d), (g), (j) are $XT$ plots with a 2 mm window, $kf$ spectrum and GOF respectively computed for this window.

Download Full Size | PDF

The bottom and right boundaries were set to be absorbing layers to minimize reflections. Since the acoustic impedance was held constant over the whole domain (${c_L}$ and $\rho $ constant), the longitudinal wave did not reflect at different interfaces and its absorption was optimal at the right and bottom boundaries. This restricted the analysis to propagating shear waves. The domain was discretized with linear finite elements on a regular rectangular grid with a grid spacing of $\,5\,\mu $m for $h = [{0.25,\; 0.5,\; 1} ]$ mm and 2.5 $\mu $m for $h = 0.125\; $mm. Because under-integrated linear elements are prone to spurious solutions, we applied Belytchko-Bindeman strain hourglass suppression [7]. The equations were integrated in time using an explicit time-stepping method to generate the full vibration velocity vector (time derivative of the dynamic displacement vector) at each space and time point. Only the vertical component of the vibration velocity was analyzed for direct comparison to experimental results (see also [8,9] for details on implementing simulations in OnScale).

Along the top boundary of the domain ($z$ = 0), we applied a spatially- and temporally-varying pressure load to simulate the acoustic push applied to the surface by the A$\mu $T transducer in our experiments [1]. A line source was considered to mimic experimental push conditions [911]. The acoustic intensity $I(x )$ of the push across the line source was considered Gaussian with a full-width-at-half max (FWHM) of 500 $\mu $m. We used this Gaussian model to define the spatial push profile:

$$P(x )= \frac{{({1 + {R^2}} )I(x )}}{{{c_{air}}}} \approx \frac{2}{{{c_{air}}}} \cdot {P_0} \cdot \exp \left[ { - 4 \cdot \log 2 \cdot {{\left( {\frac{x}{d}} \right)}^2}} \right],$$
where R and ${c_{air}}$ are the reflection coefficient and speed of sound in air, respectively ($R$ ≈ 1, ${c_{air}}$ = 340 m/s), ${P_0}$ = 5 kPa is the amplitude of the applied pressure, and $d$ = 500 $\mu $m is the push width, defined here as the FWHM of the acoustic intensity.

We simulated two different temporal excitation profiles. First, the temporal push profile was very short, i.e. shorter than the time required for a shear wave to propagate across the excitation zone. This satisfied the pressure confinement condition for excitation and launched ultra-broadband pulses of mechanical waves. In the second situation, the envelope of the temporal push was modulated by a harmonic law with a 1 kHz repetition rate.

2.1.1. Broadband excitation

For broadband excitation, the push temporal profile followed a super-Gaussian function

$$s(t )= \exp \left( { - 16 \cdot \log 2 \cdot {{\left( {\frac{{t - {t_0}}}{T}} \right)}^4}} \right),$$
where T was the push duration, defined as the full-width-at-half-maximum of the envelope function $s(t )$. A time delay ${t_0}$ was introduced to avoid impulsive loading at time $t = 0$. The push profile was centered, and full symmetry was assumed at the left boundary. To fulfill the pressure confinement condition for mechanical wave excitation, the duration of the push was considered short and equal to $T = \; 100$ $\mu $s.

Wave behavior in bounded media can be complicated and result in multiple guided modes (Fig. 1(b)). Using the group velocity of propagating waves to evaluate the shear modulus can produce serious mistakes [2]. To accurately reconstruct the shear modulus from guided wave propagation, guided modes must be sorted. A common method for mode sorting computes the two-dimensional (2D) Fourier Transform of the space-time profile (or $XT$ plot, see Fig. 1(b)) to obtain its wavenumber-frequency spectrum or ($kf$ plot, see Fig. 1(e)). In the case of guided broadband propagation, the $kf$ plot clearly shows the two first modes of guided behavior (${A_0}$ and ${S_0}$ modes).

The spatial range used in the computation of the $kf$ spectrum (window shown as the dashed rectangle in Figs. 1(b, c, d)) defines the lateral spatial resolution in the reconstructed modulus. We used rectangular windows smoothed with a Gaussian function at both edges:

$$W({x,t} )= \; {w_x}.\; {w_t},$$
where
$${w_x} = \; \left\{ {\begin{array}{c} {1\; if\; {x_{min}} \lt x \lt \; {x_{max}}}\\ {\exp ( - \frac{1}{2}{{\left( {4\frac{{x\; - \; {x_{max}}}}{{{\sigma _x}}}} \right)}^2})\; if\; x \gt {x_{max}}}\\ {\exp (\frac{1}{2}{{\left( {4\frac{{x\; - \; {x_{min}}}}{{{\sigma _x}}}} \right)}^2})\; if\; x \lt {x_{min}}}\\ {{x_{min}} = {x_{center}} - \frac{{{L_{win}}}}{2} + {\sigma _x}}\\ {{x_{max}} = {x_{center}} + \frac{{{L_{win}}}}{2} - {\sigma _x}}\\ {{\sigma _x} = \; \frac{{{L_{win}}}}{{10}}} \end{array}} \right. \textrm{and}\,{w_t} = \left\{ {\begin{array}{c} {1\; if\; {T_{min}} \lt t \lt \; {T_{max}}}\\ {\exp ( - \frac{1}{2}{{\left( {4\frac{{t\; - \; {T_{max}}}}{{{\sigma_t}}}} \right)}^2})\; \; if\; t \gt {T_{max}}}\\ {\exp (\frac{1}{2}{{\left( {4\frac{{t\; - \; {T_{min}}}}{{{\sigma_t}}}} \right)}^2})\; \; if\; t \lt {T_{min}}}\\ {{T_{max}} = {L_t} - \; {\sigma_t}}\\ {{T_{min}} = {\sigma_t}}\\ {{\sigma_t} = \frac{{{L_t}}}{{10}}} \end{array}} \right.$$
with ${L_t}$ the total recording time, and ${L_{win}}$ and ${x_{center}}$, respectively, the size and position of the window in the $x$-dimension.

Assuming appropriate temporal sampling of the propagating elastic wave, a relatively large window in space results in a localized $kf$ spectrum where guided wave modes are clearly defined. When guided modes are sampled sufficiently, shear moduli can accurately be reconstructed using an inverse method where the dispersive wave modes are calculated for a range of stiffness values (at fixed thickness). Summing the energy covered by the theoretical fit to the computed dispersion curves (in $kf$ space) of ${A_0}$ and ${S_0}$ modes provides a value of elastic modulus that corresponds to the best-fit solution based on maximum coverage [9]. For example, the goodness of fit (GOF) is equal to 1 when the theoretical solution best matches the dispersion curves obtained by applying a 2D Fourier transform to the measured wavefield.

To demonstrate the influence of the spatial window on modulus reconstruction, we performed computations for a broad range of shear moduli. Figures 1(e), (f), (g) show how resolving guided modes is less accurate as the window size shrinks, resulting in a broadened spectrum with corresponding GOF changes (Figs. 1(h), (i), (j)). When the size of the window used to reconstruct the modulus is smaller than a certain value, fitting is incorrect (Fig. 1(g)), and reconstruction of the shear modulus becomes impossible (Fig. 1(j)).

Note that only the first two guided modes were used for fitting. Although an infinite number of modes can be obtained in simulations for an infinitely short in time and infinitely narrow in space excitation, localization of the push in time and space is usually limited in practice. This in turn limits the bandwidth of the launched mechanical transient and, therefore, the number of generated guided modes. Practical implementations of OCE are usually limited to ${A_0}$ and ${S_0}$ modes [12].

The goal of numerical simulations was to define the minimum window size for which the computed $kf$ spectrum can still be fitted with a theoretical solution for ${A_0}$ and ${S_0}$ modes to produce accurate values of reconstructed shear moduli. We investigated how the minimum window size depends on the A$\mu $T push characteristics and the medium thickness. When the window size was determined, we ‘moved’ the window along the medium surface through the interface separating two distinct isotropic media (i.e. with different stiffnesses) to finally define the lateral elastographic resolution in OCE of bounded structures.

2.1.2. Harmonic excitation

Quasi-harmonic excitation of guided waves was produced by low-frequency (kHz range) modulation of the high-frequency (1 MHz) push signal described by the expression below:

$$s(t )= {{\cos }^2}\left( {\frac{\omega }{2}t} \right) \times \exp \left[ { - {{\left( {\frac{{\omega ({t - {t_0}} )}}{{4\pi }}} \right)}^4}} \right],$$
where $\omega $ is the angular frequency of the modulation and ${t_0}$ a time delay introduced to avoid impulsive loading at time $t = 0$. The geometry of the guided wave excitation in the two-part medium (see Fig. 2(a)) was the same as that for the broadband excitation described in the previous section. A typical space-time profile ($XT$ plot) for a 1 kHz excitation is represented in Fig. 2(b), and a vertical cross-section (i.e. the distribution of vertical displacement over the propagation distance) for a 6 ms time moment is shown in Fig. 2(c).

 figure: Fig. 2.

Fig. 2. Numerical simulation in OnScale of guided wave propagation in a two-part nearly incompressible linear isotropic material of uniform thickness ($h = 0.5$ mm in this figure) bounded below by water and above by air (mimicking cornea boundary conditions). Quasi-harmonic signals of guided waves are launched with acoustic micro-tapping [10] from air. Shear moduli in parts I and II used in this figure are equal to $G = {G_1} = 30\; $kPa and $G = {G_2} = 60\; $kPa respectively. The interface between the two parts is at $x = 10$ mm. The push profile is considered Gaussian in space and harmonically modulated in time, as described in Eq. (4). (a) The medium geometry used in the OnScale simulation was the same as that for broadband excitation. (b) Space-time ($XT$) plot of guided wave propagation at the medium surface for a 1 kHz modulation frequency. (c) Vertical vibration velocity as a function of propagation distance at $t = 6$ ms after the excitation. Wavelengths ${\lambda _1}$ and ${\lambda _2}$ correspond to regions of part I and part II, respectively.

Download Full Size | PDF

For quasi-harmonic signals, signal processing described in the previous section is not optimal and unnecessary. Indeed, the wavelength of the propagating wave can be measured directly at every point of the medium using the distance between 2 consecutive signal maxima. As seen in Fig. 2(c), the wavelength ${\lambda _2}$ in the second part of the medium ($x \,>\, 10$ mm) is larger than that in the first part according to the difference in mechanical moduli of parts I and II. Wavenumbers in parts I and II can be calculated as reciprocals of wavelengths. The wavenumber computed at every point of the medium at the modulation frequency was then used to determine the shear modulus G along the direction of wave propagation using the solution for the ${A_0}$ mode for this geometry [12]. The dependence of G on the distance from the excitation was then investigated to determine the transition zone between medium parts, i.e. the lateral elastographic resolution.

2.2. Two-part phantom

Polyvinyl alcohol (PVA) was used to make a two-component sample with controllable mechanical properties. Optical properties of the phantom were tuned by introducing titanium dioxide micro-particles. The phantom was created using protocols adapted from [13] and followed the recipe of [1], with a few small modifications. Specifically, two separate concentrated solutions of TiO2 (0.1% wt and 0.075% wt) were prepared in water and dispersed using a magnetic stirrer. 12% wt and 8% wt PVA (146-186 kDa, > 99% hydrolyzed, CAS: 9002-89-5, Sigma-Aldrich) was then added to the corresponding TiO2 nanoparticle solutions. The solutions were then covered and stirred on a hot plate at 140°C for approximately two hours to dissolve PVA in solution. Once PVA was fully dissolved, the 12% PVA (.075% TiO2) solution was poured into a circular mold. Pressure was then applied to the surface of the phantom using a glass coverslip to achieve a relatively homogenous, flattened, thin shape and the material was allowed to cool to room temperature (35°C). Once at room temperature, a sharp blade was used to cut the phantom in half and part of the PVA was removed.

Molten 8% PVA (.1% TiO2) was then poured into the mold and a coverslip applied across both parts to create a 2-part boundary with matched thickness of ∼0.5 mm. The circular mold had a diameter of 10 cm. Both solutions were degassed using a vacuum chamber prior to pouring into the mold. Once in the mold, the phantom underwent three freeze-thaw cycles (cooled to -20°C then warmed to 35°C) before it was placed in water for up to 12 hours prior to imaging to prevent dehydration. The two-part phantom was then clamped at the edge and placed on top of a water bath (Fig. 3(a)) to create boundary conditions that result in total internal reflection of elastic waves (air on top and water on bottom, similar to that of the cornea).

 figure: Fig. 3.

Fig. 3. a) The surface of a 2-part PVA phantom with welded boundary, clamped, and suspended on water. The higher concentration of nanoparticles (.1% TiO2) can be seen in the 8% PVA, where the lower concentration of nanoparticles (.075% TiO2) made the 12% PVA more transparent. The red dotted line denotes the approximate OCT B-scan location (not to scale). b) OCT structural image (B-scan) of the 2-part welded phantom. The soft 8% PVA material on the left of the white dotted line appeared brighter due to higher optical scattering; the hard 12% PVA is on the right. The OCT display range was 90-130 dB and a non-uniform aspect ratio was used for visualization.

Download Full Size | PDF

2.3. A${\boldsymbol \mu }$T-OCE system

A cylindrically focused, resonant (1.05 MHz central frequency) air-coupled A$\mu $T transducer was directed at the phantom surface to induce a localized displacement within a thin strip area (approximately 11 mm long and 0.6 mm wide) via reflection-based acoustic radiation force. A temporally short (100 $\mu $s), chirped excitation was used to provide approximate pressure confinement in the excitation region. Displacement relaxation resulted in the excitation of quasi-planar broadband pulses of mechanical (both surface and oblique bulk shear) waves of about 0.6 mm duration (i.e. the same as the source width). Detailed information about the A$\mu $T transducer and its operating principle can be found in [1,10,14]. Due to the wave-reflective interface at the phantom bottom (phantom-water interface), guided waves developed as the elastic wave propagated away from the excitation source. As such, accurate modulus reconstruction required the analysis of guided waves conducted at longer (multiple wavelength) distances from the excitation source.

Elastic waves were tracked in the phantom material using a spectral domain phase-sensitive OCT (PhS-OCT) system operating in M-B mode. Details and schematics of the system can be found in the Supplementary material of a previous publication [11]. A single A$\mu $T-OCE scan generated and tracked elastic waves (see details in [11,15]). The A-line rate of the system (temporal resolution), determined by the sampling rate of the 1024-pixel line-scan InGaAs array, was 90 kHz, sufficient for alias-free acquisition of induced mechanical waves inside the phantom. The spatial resolution was ∼ 10 $\mu $m axially and ∼ 15 $\mu $m laterally. 512 repeated OCT A-scans (referred to as an M-scan) were taken at a set of distances from the A$\mu $T focus. Propagating guided waves were tracked with OCT along the surface using a sequence of M-scans conducted at 256 spatial locations over a 10 mm range ($dx = 54.7$ $\mu $m) for every time instant relative to the start of each sequence. The entire spatio-temporal scan with A$\mu $T excitations took approximately 1.5 seconds.

The M-B scan dataset was then used to track the propagating wave disturbance based on local vibration velocity. The axial (or vertical) vibration velocity inside the sample at a given location ($\Delta {v_z}({x,z,t} ))$ was obtained using the optical phase difference $\mathrm{\Delta }{\varphi _{opt}}({x,z,t} )$ between two consecutive A-line scans at each location using the following equation [16]:

$$\Delta {v_z}({x,z,t} )= \; \frac{{\mathrm{\Delta }{\varphi _{opt}}({x,z,t} )\bar{\lambda }}}{{4\pi \bar{n}f_s^{ - 1}}},$$
where $\bar{\lambda }\,$ was the center wavelength of the broadband light source, $\bar{n}$ was the refractive index of the medium, and ${f_s}$ was the sampling frequency. The refractive index ($\bar{n})$ was assumed to be 1.3 and the differences due to TiO2 and PVA concentrations were considered negligible.

3. Results

3.1 Results of numerical simulations

Exploring lateral resolution in OCE for bounded media requires phantoms with predefined mechanical and geometrical properties as well as a tunable OCE system to consider a broad range of excitation parameters, which is not easy to do experimentally. FEM simulations performed in Onscale, however, can closely replicate experimental situations. Such simulations were used to greatly simplify the parametric study of phantom thickness, its mechanical moduli, as well as excitation parameters of the OCE system.

3.1.1 Broadband excitation

For the boundary conditions of bounded media, multiple guided modes can be generated simultaneously. The appropriate modes should be sorted and analyzed to convert their propagation characteristics into mechanical moduli. Note that the group velocity of propagating guided modes cannot be converted to the material shear modulus [2,12], and thus different methods must be used. As described above, a common method uses the 2D Fourier Transform of recorded $XT$ plots (Figs. 1(b), (c), (d)) to compute $kf$ spectra (Figs. 1(e), (f), (g)). The $kf$ spectra can then be used to identify propagating guided modes. Fitting the computed $kf$ spectrum with a theoretical solution of guided wave propagation can determine the material shear modulus G (see Figs. 1(e), (f), (g)). The GOF, defining the best estimate of the modulus from the fit, depends on the spatial window size used to compute the $kf$ spectrum. The minimum window size is defined by the threshold where the shear modulus G can still be accurately computed.

In FEM simulations, ${G_1}$ (in part I) was held constant at $30\; $kPa and ${G_2}$ (in part II) was varied from 30 to 120 kPa (corresponding to bulk shear wave speeds ${c_s}$ of $5.47$ m/s and $10.95$ m/s, respectively). Figure 4 shows the reconstructed shear modulus as a function of the window size for different values of ${G_2}$ (curves of different colors) and for different thicknesses of the bounded material (different panels of Fig. 4). All reconstructions were performed at the center of the second part ($x = 15$ mm). It is important to note that when guided waves pass the interface, the spatial duration of the pulse (or mean wavelength) changes because of the change in shear wave speed. For harmonic excitation, this results in a wavelength change (frequency is conserved). Interestingly, the results show that the minimum window size required to accurately reconstruct the elastic modulus does not depend on the shear modulus ${G_2}$ in part II of the medium. In other words, the minimum window size for shear modulus reconstruction does not depend on the wavelength of propagating guided waves.

 figure: Fig. 4.

Fig. 4. Window-size dependence of the reconstructed shear modulus G for the values shown in the insert to panel (a) at different layer thicknesses: (a) $h = 0.25$ mm, (b) $h = 0.5$ mm, (c) $h = 1$ mm.

Download Full Size | PDF

The independence of window size on wavelength may appear to contradict previous results (see [1]) where surface (or Rayleigh) waves were used and the elastographic resolution was determined by the wavelength. There is no contradiction, however, because the ultimate resolution in dynamic OCE is still defined by the mechanical signal wavelength, but guided waves add additional constraints. Importantly, bounded materials require a different mechanical model than surface (Rayleigh) waves that were previously used to reconstruct shear moduli. For the case of a bounded material, sufficient propagation distance is required to form guided modes, which cannot happen over sub-wavelength propagation distances. Correspondingly, Fig. 4 shows that the minimum window size to accurately reconstruct modulus depends on the layer thickness but remains greater than the pulse excitation width.

To explain the ‘paradox’ presented in Fig. 4, we must first understand acoustic wave generation with A$\mu $T. Impulsive acoustic radiation force applied to the sample surface creates two surface waves, Rayleigh (or SAW) and super-shear evanescent wave (or SEW waves) [17], and a bulk shear wave propagating at an angle to the surface. In general, the propagation angle of the oblique shear wave depends on the ratio of shear and bulk moduli in the medium [18], but because ${G_{1,2}} \ll K$ ($K$ is the bulk modulus), the directivity pattern for the oblique shear wave does not depend on the shear modulus G and has a maximum at about $\vartheta \approx 26^\circ $ to the sample normal. This explains why the minimum window size is nearly wavelength independent and why it mainly depends on thickness. Indeed, the propagating oblique shear wave in the layer and its conversion to SEW [17] at each reflection from layer interfaces creates guided waves in the medium. When $\vartheta $ is fixed for all values of G, the distance at which guided waves are fully formed does not depend on G. The layer thickness changes the number of reflections per unit length and, thus, changes the minimum distance at which guided waves are formed, or the minimum window size within which guided modes can be identified.

Once the minimum window size needed to identify a guided wave is known, the transition of guided modes through the interface ($x = 10$ mm) can be explored. By translating the specified window through the entire medium length ($0\,<\, x \,<\, 20$ mm, see Fig. 5(a)), the reconstructed value of the shear modulus is mapped (Fig. 5(b)). The modulus change through the transition zone was fit to a combination of two sigmoid functions corresponding to the regions before and after the interface, respectively:

$${G_{fit}} = \frac{a}{{1 + {e^{ - b({x - {x_0}} )}}}} + c$$
where a, b and c are fitting parameters, and ${x_0} = 10\; $mm corresponds to the interface between parts.

 figure: Fig. 5.

Fig. 5. (a) Schematics of moving window, defined in Fig. 4, across the interface between two medium parts. (b) Dependence of the reconstructed shear modulus G on the distance from the source (black dots), its fit with a sigmoid function (Eq.(6), blue curve) and derivative of the sigmoid function (red dashed curve). Black horizontal bar shows the computed transition zone (or lateral elastographic resolution) at 6 dB (full width half maximum, FWHM). Shear moduli of two medium parts used in this figure are ${G_1} = 30$ kPa and ${G_2} = 60$ kPa; medium thickness is $h = 0.5$ mm. The interface between part I and part II corresponds to $x = 10$ mm.

Download Full Size | PDF

The derivative of this function can be used to quantify the transition zone length, i.e. determine the lateral spatial resolution of dynamic OCE for bounded media.

The relationship between the window size and resolution for materials with different thickness was explored to highlight subtle differences between processing artifact and fundamental limits. Specifically, the ‘optimal’ window size to reconstruct the modulus can be determined automatically by varying its size and applying it to $XT$ plot processing, as shown in Fig. 6. When the window size was above the optimal one, the transition zone reduced proportional to the window decrease. However, there was a limit where further decrease of the window size did not change the transition zone and, at that, reduced the accuracy of the modulus reconstruction. This also provided a practical method to determine the minimum window size to optimize accuracy without reducing resolution; the optimal window can be taken from the truncation point. The independence of the window size on G is very important because the same window can be used for any gradients of mechanical properties in a bounded medium with constant thickness.

 figure: Fig. 6.

Fig. 6. Dependence of the transition zone on window size at different layer thicknesses: (a) $h = 0.25$ mm, (b) $h = 0.5$ mm, (c) $h = 1$ mm. Shear modulus of part I of the medium used in this figure is ${G_1} = 30$ kPa. The ratio of moduli for the two parts is shown in the insert. Dashed red line corresponds to the optimal window size.

Download Full Size | PDF

As shown in Figs. 6(a)-(c), the optimal window size (or the truncation point in the dependence of the transition zone on window size) strongly depends on layer thickness. Figure 7 shows how the transition zone changed with window size at different layer thicknesses. The relationship between spatial resolution and thickness was obtained for a broad range of ${G_2}$, and the empirical dependence of spatial resolution on thickness was defined with a linear function:

$$\Delta x(h )= 2.05 \times h + 0.32\; \textrm{mm},$$
where the transition zone decreased with thickness h (taken in mm in Eq. (7)) almost linearly over the range of layer thicknesses investigated. This transition zone equals the elastographic spatial resolution in wave-based OCE using guided mechanical waves. Note that in dynamic OCE, OCT is used to track propagating mechanical waves. Thus, the layer thickness can be measured simultaneously and, therefore, the window size can always be properly chosen using the dependences shown in Figs. 6 and 7.

 figure: Fig. 7.

Fig. 7. Dependence of the transition zone width on the thickness of a two-part medium. Shear modulus of the first part of the medium used in this figure is ${G_1} = 30$ kPa; in the second part shear moduli varied over the range ${G_2} = $ (33-120) kPa. All depicted values correspond to mean values obtained for different values of ${G_2}$; the error bars correspond to their variations. Dashed red line corresponds to fitting the points with a linear function (Eq. (7)).

Download Full Size | PDF

3.1.2 Harmonic excitation

When a harmonic excitation is employed, the frequency of propagating waves is always fixed, but the wavelength (or wavenumber) changes from point to point depending on the local thickness and mechanical modulus. As we pointed out in the Methods section (see Fig. 2), measuring the local wavenumber at a known sample thickness can be used to reconstruct the mechanical modulus ${G_{1,2}}$ using the ${A_0}$ mode dispersion relation [12]. Using a window to fit the guided wave spectrum for harmonic excitation is meaningless because the wavenumber can be directly determined from the space-time wavefield (see Fig. 2(c)).

Figure 8 shows the dependence of the lateral elastographic resolution on the carrier frequency of quasi-harmonic excitation for different thicknesses of the two-part medium. First, the computed transition zone, which corresponds to the spatial resolution, depends on the layer thickness in a similar fashion to that illustrated in Fig. 7 for broadband excitation. Second, Figs. 8(a), (d), (g) demonstrate that the transition zone is also frequency dependent. The reason for this is the frequency-dependent wavenumber of the guided waves. The resolution can therefore be improved by increasing the excitation frequency. However, there is a problem in using this strategy. As Figs. 8(f), (h), (i), show, increasing the excitation frequency leads to fluctuations in the reconstructed moduli which become significant in the example shown in Fig. 8(i). These fluctuations appear for conditions (frequency and thickness) where both ${A_0}$ and ${S_0}$ modes, propagating with different speeds, coexist creating phase disturbances in the wavefield. Determining the proper frequency for OCE studies for quasi-harmonic excitation can be extremely difficult when the medium properties are unknown, especially when it is heterogeneous with varying mechanical properties and thickness.

 figure: Fig. 8.

Fig. 8. Dependence of the reconstructed shear modulus G on the distance from the source (black dots) in the two-part medium with shear moduli ${G_1} = 30$ kPa and ${G_2} = 60$ kPa, respectively, for the case of quasi-harmonic excitation. The excitation frequency and medium thickness are shown on the left and on the top of the panels, respectively. The interface between part I and part II corresponds to $x = 10$ mm. Fits of the transition zone with the sigmoid function (Eq. (6)) are plotted with blue curves, and the derivative of the sigmoid function is plotted with the red dashed curve in every panel. The black horizontal bar shows the computed transition zone (or lateral elastographic resolution). The lateral resolution in panel (i) is undetermined due to high fluctuations in reconstructed moduli.

Download Full Size | PDF

3.2 Results of AμT-OCE experiments

To confirm the findings of numerical simulations, A$\mu $T-OCE experiments were performed in the two-part phantom described in the Methods section. First, wavefields ($XT$ plots) were generated and recorded in each part of the phantom far (∼1 cm) from the interface between them. The $XT$ plots in the sections containing only 8% PVA and 12% PVA are presented in Figs. 9(a), (c) respectively. The elastic modulus was reconstructed in both parts using the $kf$ spectra (Figs. 9(b), (d)) computed from the recorded guided wavefields (${G_1} = ({28 \pm 2} )$ kPa and ${G_2} = ({62 \pm 4} )$ kPa, corresponding to bulk shear wave speeds ${c_s}$$= 5.29\; $m/s and $7.87\; $m/s, respectively). The values and error bars (standard deviation) were computed using fits on 5 different scans with a 10 mm window size in each part of the phantom. Then, the optimal window size was determined as the minimum range where the reconstruction did not change modulus values. A 4 mm optimal window (Fig. 9(e)) was found by decreasing the window size in the $XT$ plots. Note that this value was close to that predicted in numerical simulations for a 0.5 mm material. As also demonstrated in numerical simulations, the optimal window size was obtained independent of the material modulus, i.e. regardless of the elastic modulus (Fig. 9(e)).

 figure: Fig. 9.

Fig. 9. Space-time waveforms (XT plots) recorded experimentally across 6 mm in a two-part phantom (see Fig. 3) in (a) part I (${G_1} = ({28 \pm 2} )$ kPa) and (c) part II (${G_2} = ({62 \pm 4} )$ kPa) and their corresponding $kf$ spectra (b) and (d), respectively. (e) Dependence of the reconstructed shear modulus on window size in each part of the phantom. (f) Dependence of mechanical modulus G on the distance from the A$\mu $T source in the direction from part II to part I across the interface. Fit of the transition zone with the sigmoid function (Eq. (6)) is plotted with a blue curve, and the derivative of the sigmoid function is plotted with the red dashed curve. The black horizontal bar shows the computed transition zone (or lateral elastographic resolution).

Download Full Size | PDF

Finally, propagation of guided waves was investigated in the direction perpendicular to the interface between the medium parts. The optimal window size (4 mm) was used to compute the medium mechanical properties across the transition between medium parts. Figure 9(f) demonstrates that an experimentally obtained transition zone is about $\Delta x \cong \; $1.1 mm (similar to $\Delta x \cong \; $1.3 mm obtained in numerical simulations shown in Fig. 7).

4. Discussion

In this study, we explored the lateral elastographic resolution of wave-based OCE in bounded soft tissue. Contrary to the broadly accepted opinion that the resolution is fully defined by the capabilities of the imaging system, we demonstrate here that it is not always true. Clearly, geometrical constraints and the appropriate mechanical model play an important role in the ultimate resolution of dynamic OCE.

In Ref. [1] we demonstrated that for Rayleigh waves, the excitation of waves propagating along the interface between materials with different stiffness changes the propagating wave’s shape. This shape is complicated over the region equal to approximately the spatial width of the propagation pulse and, thus, smooths the transition between reconstructed moduli. For the case of Rayleigh waves, the best resolution is not equal to the optical resolution of the OCT imaging system. Practically, it is determined by the propagating elastic wave and can be up to one order of magnitude worse than the optical imaging resolution.

In this study, we have shown that the resolution of the imaging system is not critical for dynamic elastography of bounded media. Thus, we expect comparable elastographic resolution in OCE to that in USE, with much worse resolution compared to OCT structural imaging. However, there are numerous advantages to OCT over ultrasound as a host imaging technique. For example, OCE can provide co-aligned tomographic images (based on light scattering) as well as functional mapping of vascularization and local fiber orientation [19]. This can provide additional diagnostic information as well as detailed geometric information of the medium under study (collagen fiber orientation, boundary conditions, etc.) to determine the most appropriate material model to analyze the propagating wave. For example, a precise measurement of layer thickness is critical to reconstruct shear moduli from propagating guided waves. Additionally, micro-tapping OCE can launch elastic waves in a medium without contact. Thus, a coupling medium is not needed, and the measurement system does not affect the propagation medium (no change in boundary conditions and no induced prestress), enabling measurement of sensitive tissue such as cornea, burn wounds, grafts and ulcers.

Regardless of the host imaging method, an additional interface, i.e. bounding the tissue, further reduces elastographic resolution. Guided wave behavior is stimulated in the bounded layer, producing one or multiple dispersive guided modes with propagation speed directly related to frequency and thickness. The group velocity of guided modes cannot be as easily converted to the shear modulus as that for bulk shear waves, and modulus reconstruction from guided wavefields should be performed with care. Guided modes are formed mainly by multiple reflections of the oblique bulk shear wave inside the layer with its conversion to super-shear waves and, therefore, requires a certain distance from the source or stiffness interface for guided modes to form. The propagation speed within the transition zone is variable, and modulus reconstruction within it is inaccurate.

In general, the directivity pattern of the oblique shear wave depends on material properties. For aluminum and brass, for example, the propagation angle will differ [18]. However, in a nearly incompressible medium, i.e. in soft tissue, the shear modulus is typically five orders of magnitude smaller than the bulk modulus. This means that the angle of bulk shear wave propagation does not depend on shear modulus [18]. The distance from the source at which the bulk shear wave hits the front surface after its reflection from the bottom of the layer is also relatively independent of the modulus under the ray approximation. Thus, the near field for guided behavior is roughly independent of the wavelength but depends on the medium thickness. Whether broadband pulses or harmonic mechanical waves are employed, and regardless of local mechanical moduli, the elastographic resolution in dynamic OCE of bounded media is mainly determined by the medium thickness.

The relationship between resolution and thickness was demonstrated directly for the case where the thickness was larger than half of the wavelength (or the spatial pulse width for broadband excitation, see Eq. (7)). The limiting case for decreasing thickness was not explored. In other words, the question remains: how does the transition zone behave when the layer thickness tends to zero? Unfortunately, Fig. 7 does not answer this question because the minimum layer thickness we could achieve in numerical simulations was $h = 125$ $\mu $m. Further thickness reduction led to either numerical dispersion in the solution or further grid refinement (less than 2.5 $\mu $m) keeping all other dimensions in the numerical phantom, which resulted in numerical instabilities. However, additional simulations were performed for a different excitation width ($d = 1$ mm) for layer thicknesses varying over the (0.125–1) mm range. The minimum transition zone (or the highest lateral elastographic resolution) we could reach in both series was equivalent to the excitation width. Thus, when the thickness becomes much smaller than the mechanical pulse excitation width, the transition zone reaches a minimum of the excitation width, i.e. the mechanical pulse spatial extent. This conclusion also follows from the 0.32 mm offset in Eq.(7).

The linear dependence observed in Fig. 7 is only valid in the situation where guided waves propagate, since it is linked to the XT window size necessary to capture this behavior. As thickness increases, the guided behavior detected on the top surface will gradually give way to a Rayleigh surface wave, as expected for semi-infinite media. This connects our previous results obtained with Rayleigh waves with the results of the current study – the best mechanical resolution that can be obtained in dynamic OCE using broadband excitation is on the order of the excitation pulse width.

An important corollary of this work, and the previous exploration using Rayleigh waves, is that the ultimate resolution in dynamic OCE depends on the assumed (or most appropriate) mechanical model and the reconstruction method required to achieve quantitatively accurate contrast. Others [20], for example, showed that the ultimate resolution in a bulk material can be a few times better than half of the wavelength when artifacts induced by the interface are suppressed. For guided waves, however, we do not see a clear way to “unmix” guided modes or analyze them in the near field, accounting for highly variable experimental conditions in real tissue. One possible way to avoid guided waves for modulus reconstruction in bounded media is to use the SEW [17] propagating in the near field along the tissue surface. The wave speed of the SEW is almost twice that of the bulk shear wave. This fast propagation speed suggests that it can be separated from the main guided waves. Unfortunately, the SEW attenuates very quickly, which limits the imaging range or requires multiple pushes over the imaging area.

In this work we considered both broadband and harmonic excitations since both methods are typical for dynamic OCE [11,2124]. Although harmonic signals generally have better signal-to-noise ratio than broadband pulses, harmonic waves should be treated with care. When a single frequency is generated and material properties are unknown, computing the local wavenumber can produce serious error due to the presence of multiple guided modes. Here we calculated the local wavenumber assuming that both parts of the phantom propagated only the ${A_0}$ mode. This, however, is not a given. For the example shown in Fig. 1(e), the ${S_0}$ mode had much higher energy at 4 kHz compared to the ${A_0}$ mode, whereas at 2 kHz the situation is completely opposite. Thus, if ${G_1}$ is 4 times higher than ${G_2}$, the wavenumber calculated for the same frequency in the first part of the medium for the ${A_0}$ mode will correspond to the wavenumber of the ${S_0}$ mode in the second part, and the computed modulus will be incorrect.

${A_0}$ and ${S_0}$ modes can also coexist, which leads to increased fluctuations in the reconstructed moduli in both parts of the medium (see Fig. 8(h)). Such fluctuations can totally corrupt the reconstruction (as in Fig. 8(i)). The same effect will occur when medium thickness changes. Note that while the ${A_0}$ mode tends to dominate in anisotropic tissues such as the cornea [9], mode jumping due to the finite cornea diameter and fixed boundary conditions can result in structural low frequency resonance (shape modes in the x-direction) that can be difficult to recognize for a fixed frequency. Therefore, harmonic waves at different frequencies should be analyzed in practice to make sure that there are no mode jumps and/or mode competition. Separating reconstruction artifacts from true material heterogeneities may be difficult.

In addition to the local wavenumber calculation, we also explored phase velocity computation, as proposed in Refs. [23,25]. Results for this method can be seen in the Supplemental document. Although there appeared to be a certain relationship between medium properties, excitation frequency and thickness, where the spatial resolution appeared to reach the wavelength limit (Suppl. Fig. 1(d)), it can be extremely difficult to reliably reconstruct the modulus using a single frequency in general due to enormous phase instabilities in the guided wavefield (see Suppl. Figures 1(e-i) and Suppl. Fig. 2(b)), except for the case where the probing wavelength is much larger than the layer thickness (Suppl. Fig. 2(a)). However, very low frequency measurements are complicated by boundary conditions when the medium size is limited, as in cornea, for example. Reducing the frequency will also reduce reconstruction accuracy because low frequencies are highly dispersive. For simple cases where the interface between heterogeneities is sharp and known, high-order guided modes can be filtered out in $kf$ space, but for unknown borders between heterogeneities or for inclusions small compared to the wavelength, such local filtration may be difficult.

When broadband guided wave signals are used, the 2D Fourier transform can be used to reconstruct wave dispersion spectra and sort guided waves in the $kf$ domain. Fitting a theoretical solution to expected modes in the $kf$ spectrum over a broad frequency range avoids the artifacts associated with single frequency waves. In addition, such processing provides an avenue to ensure that the mechanical model for the medium is chosen correctly, as characterized by the goodness of fit (GOF).

As a final tradeoff in determining the resolution in guided wave-based OCE, the processing window should be optimized for accuracy without limiting resolution. Note that when broadband signals are used to reconstruct moduli, quantitative errors are reduced. If the reconstruction window is larger than optimal, this will not change the reconstructed modulus, but will rather smooth the transition between medium heterogeneities. Thus, using a broadband excitation to reconstruct the elastic modulus of thin materials in wave-based OCE is advantageous over a single frequency, and can provide mm-scale resolution in many types of human tissue.

We have studied the lateral elastographic resolution of wave-based OCE in bounded media. However, it is also important to discuss whether the material shear modulus (or moduli for anisotropic tissues [9,11]) can be resolved in depth. OCT can track mechanical waves in 3D, i.e. depth-resolved wave velocities can be computed in 3D as well. In-depth reconstruction of cornea elastic modulus from wave-based OCE was reported in Ref. [26]. In fact, such results are completely inconsistent with the nature of guided waves. Guided waves are aggregate waves that occupy the entire layer thickness. For materials with depth-dependent elastic properties, guided waves carry depth-averaged characteristics, and the reconstructed modulus that can be computed represents a single depth-averaged value. This problem can probably be overcome with near-field detection of high-frequency longitudinal shear waves [27] with the source moving over different lateral positions.

In human tissue, viscosity plays a role in mechanical wave propagation. However, it mainly influences the high-frequency content of the wave packet [21,28], producing a slight increase in phase velocities in the multiple-kHz region. For the range of frequencies analyzed using acoustic micro-tapping (typically below 5kHz), the impact of viscosity was considered second order relative to the real part of the elastic modulus. As such, the effect of viscosity on wave propagation can be ignored to greatly simplify modelling and theoretical analysis without significant deviation from the viscous condition. To measure viscosity in OCE, generation and tracking tools must preserve high frequencies. This is generally a difficult task using our technique and is outside the scope of this work. In addition, including viscosity in the model and FE simulations will not change the conclusions on guided wave resolution due to the nature of guided mode development and propagation of low-frequency energy. A similar window size must be used to obtain a reliable fk plot for elastic modulus reconstruction.

5. Conclusions

The ultimate elastographic resolution in dynamic OCE has been explored numerically and experimentally for the case of a thin bounded layer, which is typical for OCE applications. There is a broadly accepted opinion that the resolution in mapping of elastic moduli, i.e. elastographic resolution, is equal to that for the propagation wave speed, which is in turn determined by the capabilities of the imaging system. We previously showed that it is not true when waves propagate along the tissue surface (Rayleigh waves), and the accurate reconstruction of mechanical moduli may be corrupted by mode conversions in the region of about a wavelength surrounding the interface or heterogeneity. The situation becomes even more complicated when tissue is layered or bounded, which leads to the generation of dispersive guided modes in the layer. Guided wave propagation makes the phase of harmonic signals unstable and requires a propagation distance before guided modes are fully formed, i.e. there is a near field of the source or interface where wave propagation characteristics are different and, therefore, the reconstruction of mechanical moduli is inaccurate. We have shown here that a minimum transition zone is mainly driven by the medium thickness (equaling about twice the thickness) and is nearly independent of the propagation wavelength. Broadband pulse excitation of mechanical waves enables a robust method for modulus reconstruction with the best-achievable resolution using local fitting of wavenumber-frequency spectra with a theoretical model. Processing of harmonic signals for modulus reconstruction should be performed with care to avoid artifacts related to phase instabilities and mode conversions.

Funding

Coulter Translational Research Partnership Program; Life Sciences Discovery Fund (3292512); NIH (R01-AR077560, R01-CA170734, R01-EB016034, R01-EY024158, R01-EY026532, R01-HL093140); Department of Bioengineering at the University of Washington (DGE-1256082).

Acknowledgments

This work was supported, in part, by NIH grants R01-EY026532, R01-EY024158, R01-EB016034, R01-CA170734, R01-AR077560 and R01-HL093140, Life Sciences Discovery Fund 3292512, the Coulter Translational Research Partnership Program, an unrestricted grant from the Research to Prevent Blindness, Inc., New York, New York, and the Department of Bioengineering at the University of Washington. M. Kirby was supported by an NSF graduate fellowship (No. DGE-1256082). This material was based upon work supported by the National Science Foundation Graduate Research Fellowship Program under Grant No. DGE-1256082.

Disclosures

The authors declare no conflict of interests

Data availability

Data underlying the results presented in this paper may be obtained from the authors upon reasonable request.

Supplemental document

See Supplement 1 for supporting content.

References

1. M. A. Kirby, K. Zhou, J. J. Pitre, L. Gao, D. S. Li, I. Pelivanov, S. Song, C. Li, Z. Huang, T. T. Shen, R. K. Wang, and M. O’Donnell, “Spatial resolution in dynamic optical coherence elastography,” J. Biomed. Opt. 24(09), 1 (2019). [CrossRef]  

2. I. Pelivanov, L. Gao, J. J. Pitre, M. A. Kirby, S. Song, D. Li, T. T. Shen, R. K. Wang, and M. O’Donnell, “Does group velocity always reflect elastic modulus in shear wave elastography?” J. Biomed. Opt. 24(07), 1 (2019). [CrossRef]  

3. M. A. Slawinski, Waves and Rays in Elastic Continua, Third (World Scientific, 2010). [CrossRef]  

4. I. A. Viktorov, Rayleigh and Lamb Waves Physical Theory and Applications, First, Ultrasonic Technology (Springer New York, NY, 1967).

5. P. V. Krauklis and L. A. Molotkov, “Low-frequency Lamb waves in cylindrical and spherical layers embedded in an elastic medium,” J. Sov. Math. 3(1), 82–90 (1975). [CrossRef]  

6. Z. Han, S. R. Aglyamov, J. Li, M. Singh, S. Wang, S. Vantipalli, C. Wu, C. Liu, M. D. Twa, and K. V. Larin, “Quantitative assessment of corneal viscoelasticity using optical coherence elastography and a modified Rayleigh–Lamb equation,” J. Biomed. Opt 20(2), 2–5 (2015).

7. T. Belytschko and L. P. Bindeman, “Assumed strain stabilization of the 4-node quadrilateral with 1-point quadrature for nonlinear problems,” Comput. Methods Appl. Mech. Eng. 88(3), 311–340 (1991). [CrossRef]  

8. J. J. Pitre, “How University of Washington is Developing a Novel Medical Device with OnScale Simulation,” Case Study ([Online] https://onscale.com/wp-content/uploads/2019/08/Case-Study_Univ_of_Washington-Medical-Device.pdf.) (2020).

9. J. J. Pitre, M. A. Kirby, D. S. Li, T. T. Shen, R. K. Wang, M. O’Donnell, and I. Pelivanov, “Nearly-incompressible transverse isotropy (NITI) of cornea elasticity: model and experiments with acoustic micro-tapping OCE,” Sci. Rep. 10(1), 12983 (2020). [CrossRef]  

10. Ł. Ambroziński, S. Song, S. J. Yoon, I. Pelivanov, D. Li, L. Gao, T. T. Shen, R. K. Wang, and M. O’Donnell, “Acoustic micro-tapping for non-contact 4D imaging of tissue elasticity,” Sci. Rep. 6(1), 38967 (2016). [CrossRef]  

11. M. A. Kirby, J. J. Pitre, H.-C. Liou, D. S. Li, R. K. Wang, I. Pelivanov, M. O’Donnell, and T. T. Shen, “Delineating corneal Elastic anisotropy in a porcine model using noncontact OCT elastography and ex vivo mechanical tests,” Ophthalmol. Sci. 1(4), 100058 (2021). [CrossRef]  

12. M. A. Kirby, I. Pelivanov, S. Song, Ł. Ambrozinski, S. J. Yoon, L. Gao, D. Li, T. T. Shen, R. K. Wang, and M. O’Donnell, “Optical coherence elastography in ophthalmology,” J. Biomed. Opt. 22(12), 1 (2017). [CrossRef]  

13. S.-H. Hyon, W.-I. Cha, and Y. Ikada, “Preparation of transparent poly(vinyl alcohol) hydrogel,” Polym. Bull. 22(2), 119–122 (1989). [CrossRef]  

14. Ł. Ambroziński, I. Pelivanov, S. Song, S. J. Yoon, D. S. Li, L. Gao, T. T. Shen, R. K. Wang, and M. O’Donnell, “Air-coupled acoustic radiation force for non-contact generation of broadband mechanical waves in soft media,” Appl. Phys. Lett. 109(4), 043701 (2016). [CrossRef]  

15. S. Song, Z. Huang, T.-M. Nguyen, E. Y. Wong, B. Arnal, M. O’Donnell, and R. K. Wang, “Shear modulus imaging by direct visualization of propagating shear waves with phase-sensitive optical coherence tomography,” J. Biomed. Opt. 18(12), 121509 (2013).

16. R. K. Wang, Z. Ma, and S. J. Kirkpatrick, “Tissue Doppler optical coherence elastography for real time strain rate and strain mapping of soft tissue,” Appl. Phys. Lett. 89(14), 144103 (2006). [CrossRef]  

17. J. J. Pitre, M. A. Kirby, L. Gao, D. S. Li, T. Shen, R. K. Wang, M. O’Donnell, and I. Pelivanov, “Super-shear evanescent waves for non-contact elastography of soft tissues,” Appl. Phys. Lett. 115(8), 083701 (2019). [CrossRef]  

18. J. R. Bernstein and J. B. Spicer, “Line source representation for laser-generated ultrasound in aluminum,” J. Acoust. Soc. Am. 107(3), 1352–1357 (2000). [CrossRef]  

19. M. A. Kirby, P. Tang, H.-C. Liou, M. Kuriakose, J. J. Pitre, T. N. Pham, R. E. Ettinger, R. K. Wang, M. O’Donnell, and I. Pelivanov, “Probing elastic anisotropy of human skin in vivo with light using non-contact acoustic micro-tapping OCE and polarization sensitive OCT,” Sci. Rep. 12(1), 3963 (2022). [CrossRef]  

20. C. Zemzemi, A. Zorgani, L. Daunizeau, S. Belabhar, R. Souchon, and S. Catheline, “Super-resolution limit of shear-wave elastography,” EPL 129(3), 34002 (2020). [CrossRef]  

21. A. Ramier, A. M. Eltony, Y. Chen, F. Clouser, J. S. Birkenfeld, A. Watts, and S.-H. Yun, “In vivo measurement of shear modulus of the human cornea using optical coherence elastography,” Sci. Rep. 10(1), 17366 (2020). [CrossRef]  

22. F. Zvietcovich, A. Nair, Y. S. Ambekar, M. Singh, S. R. Aglyamov, M. D. Twa, and K. V. Larin, “Confocal air-coupled ultrasonic optical coherence elastography probe for quantitative biomechanics,” Opt. Lett. 45(23), 6567 (2020). [CrossRef]  

23. F. Zvietcovich, A. Nair, M. Singh, S. R. Aglyamov, M. D. Twa, and K. V. Larin, “In vivo assessment of corneal biomechanics under a localized cross-linking treatment using confocal air-coupled optical coherence elastography,” Biomed. Opt. Express 13(5), 2644–2654 (2022). [CrossRef]  

24. F. Zvietcovich and K. V. Larin, “Wave-based optical coherence elastography: the 10-year perspective,” Prog. Biomed. Eng. 4(1), 012007 (2022). [CrossRef]  

25. F. Zvietcovich, J. P. Rolland, J. Yao, P. Meemon, and K. J. Parker, “Comparative study of shear wave-based elastography techniques in optical coherence tomography,” J. Biomed. Opt. 22(3), 035010 (2017). [CrossRef]  

26. S. Wang and K. V. Larin, “Noncontact depth-resolved micro-scale optical coherence elastography of the cornea,” Biomed. Opt. Express 5(11), 3807–3821 (2014). [CrossRef]  

27. F. Zvietcovich, G. R. Ge, H. Mestre, M. Giannetto, M. Nedergaard, J. P. Rolland, and K. J. Parker, “Longitudinal shear waves for elastic characterization of tissues in optical coherence elastography,” Biomed. Opt. Express 10(7), 3699–3821 (2019). [CrossRef]  

28. A. Ramier, B. Tavakol, and S.-H. Yun, “Measuring mechanical wave speed, dispersion, and viscoelastic modulus of the cornea using optical coherence elastography,” Opt. Express 27(12), 16635–16649 (2019). [CrossRef]  

Supplementary Material (1)

NameDescription
Supplement 1       Supplemental Material

Data availability

Data underlying the results presented in this paper may be obtained from the authors upon reasonable request.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (9)

Fig. 1.
Fig. 1. Numerical simulation in OnScale of guided wave propagation in a two-part nearly incompressible linear isotropic material of uniform thickness ($h = 0.5$ mm in this figure) bounded below by water and above by air (mimicking cornea boundary conditions). Shear moduli in parts I and II used in this figure are equal to $G = {G_1} = 30\; $kPa and $G = {G_2} = 60\; $kPa respectively; medium thickness is uniform over the medium and equal to $h = 0.5$ mm. The interface between the two parts is at $x = 10$ mm. Broadband pulses of guided waves were launched with acoustic micro-tapping [10] from air. The push profile was considered Gaussian in space and super-Gaussian in time, as described in Eq. (1). (a) Medium geometry used in the OnScale simulations. A 1-mm layer of water bounds the media on the right-hand side to better match absorbing boundary conditions. (b) Space-time ($XT$) plot of guided wave propagation at the medium surface. (e) Wavenumber-frequency ($kf$) spectrum obtained from $XT$ plot (b) using an 8 mm window in part II with solutions for ${A_0}$ and ${S_0}$ modes best fitting the spectrum. (h) Goodness of fit (GOF) of the simulated spectrum as a function of ${G_1}$ describing the method of shear modulus reconstruction from the $kf$ spectrum. (c), (f), (i) are $XT$ plots with a 4 mm window, $kf$ spectrum and GOF respectively computed for this window. (d), (g), (j) are $XT$ plots with a 2 mm window, $kf$ spectrum and GOF respectively computed for this window.
Fig. 2.
Fig. 2. Numerical simulation in OnScale of guided wave propagation in a two-part nearly incompressible linear isotropic material of uniform thickness ($h = 0.5$ mm in this figure) bounded below by water and above by air (mimicking cornea boundary conditions). Quasi-harmonic signals of guided waves are launched with acoustic micro-tapping [10] from air. Shear moduli in parts I and II used in this figure are equal to $G = {G_1} = 30\; $kPa and $G = {G_2} = 60\; $kPa respectively. The interface between the two parts is at $x = 10$ mm. The push profile is considered Gaussian in space and harmonically modulated in time, as described in Eq. (4). (a) The medium geometry used in the OnScale simulation was the same as that for broadband excitation. (b) Space-time ($XT$) plot of guided wave propagation at the medium surface for a 1 kHz modulation frequency. (c) Vertical vibration velocity as a function of propagation distance at $t = 6$ ms after the excitation. Wavelengths ${\lambda _1}$ and ${\lambda _2}$ correspond to regions of part I and part II, respectively.
Fig. 3.
Fig. 3. a) The surface of a 2-part PVA phantom with welded boundary, clamped, and suspended on water. The higher concentration of nanoparticles (.1% TiO2) can be seen in the 8% PVA, where the lower concentration of nanoparticles (.075% TiO2) made the 12% PVA more transparent. The red dotted line denotes the approximate OCT B-scan location (not to scale). b) OCT structural image (B-scan) of the 2-part welded phantom. The soft 8% PVA material on the left of the white dotted line appeared brighter due to higher optical scattering; the hard 12% PVA is on the right. The OCT display range was 90-130 dB and a non-uniform aspect ratio was used for visualization.
Fig. 4.
Fig. 4. Window-size dependence of the reconstructed shear modulus G for the values shown in the insert to panel (a) at different layer thicknesses: (a) $h = 0.25$ mm, (b) $h = 0.5$ mm, (c) $h = 1$ mm.
Fig. 5.
Fig. 5. (a) Schematics of moving window, defined in Fig. 4, across the interface between two medium parts. (b) Dependence of the reconstructed shear modulus G on the distance from the source (black dots), its fit with a sigmoid function (Eq.(6), blue curve) and derivative of the sigmoid function (red dashed curve). Black horizontal bar shows the computed transition zone (or lateral elastographic resolution) at 6 dB (full width half maximum, FWHM). Shear moduli of two medium parts used in this figure are ${G_1} = 30$ kPa and ${G_2} = 60$ kPa; medium thickness is $h = 0.5$ mm. The interface between part I and part II corresponds to $x = 10$ mm.
Fig. 6.
Fig. 6. Dependence of the transition zone on window size at different layer thicknesses: (a) $h = 0.25$ mm, (b) $h = 0.5$ mm, (c) $h = 1$ mm. Shear modulus of part I of the medium used in this figure is ${G_1} = 30$ kPa. The ratio of moduli for the two parts is shown in the insert. Dashed red line corresponds to the optimal window size.
Fig. 7.
Fig. 7. Dependence of the transition zone width on the thickness of a two-part medium. Shear modulus of the first part of the medium used in this figure is ${G_1} = 30$ kPa; in the second part shear moduli varied over the range ${G_2} = $ (33-120) kPa. All depicted values correspond to mean values obtained for different values of ${G_2}$; the error bars correspond to their variations. Dashed red line corresponds to fitting the points with a linear function (Eq. (7)).
Fig. 8.
Fig. 8. Dependence of the reconstructed shear modulus G on the distance from the source (black dots) in the two-part medium with shear moduli ${G_1} = 30$ kPa and ${G_2} = 60$ kPa, respectively, for the case of quasi-harmonic excitation. The excitation frequency and medium thickness are shown on the left and on the top of the panels, respectively. The interface between part I and part II corresponds to $x = 10$ mm. Fits of the transition zone with the sigmoid function (Eq. (6)) are plotted with blue curves, and the derivative of the sigmoid function is plotted with the red dashed curve in every panel. The black horizontal bar shows the computed transition zone (or lateral elastographic resolution). The lateral resolution in panel (i) is undetermined due to high fluctuations in reconstructed moduli.
Fig. 9.
Fig. 9. Space-time waveforms (XT plots) recorded experimentally across 6 mm in a two-part phantom (see Fig. 3) in (a) part I (${G_1} = ({28 \pm 2} )$ kPa) and (c) part II (${G_2} = ({62 \pm 4} )$ kPa) and their corresponding $kf$ spectra (b) and (d), respectively. (e) Dependence of the reconstructed shear modulus on window size in each part of the phantom. (f) Dependence of mechanical modulus G on the distance from the A$\mu $T source in the direction from part II to part I across the interface. Fit of the transition zone with the sigmoid function (Eq. (6)) is plotted with a blue curve, and the derivative of the sigmoid function is plotted with the red dashed curve. The black horizontal bar shows the computed transition zone (or lateral elastographic resolution).

Equations (8)

Equations on this page are rendered with MathJax. Learn more.

P ( x ) = ( 1 + R 2 ) I ( x ) c a i r 2 c a i r P 0 exp [ 4 log 2 ( x d ) 2 ] ,
s ( t ) = exp ( 16 log 2 ( t t 0 T ) 4 ) ,
W ( x , t ) = w x . w t ,
w x = { 1 i f x m i n < x < x m a x exp ( 1 2 ( 4 x x m a x σ x ) 2 ) i f x > x m a x exp ( 1 2 ( 4 x x m i n σ x ) 2 ) i f x < x m i n x m i n = x c e n t e r L w i n 2 + σ x x m a x = x c e n t e r + L w i n 2 σ x σ x = L w i n 10 and w t = { 1 i f T m i n < t < T m a x exp ( 1 2 ( 4 t T m a x σ t ) 2 ) i f t > T m a x exp ( 1 2 ( 4 t T m i n σ t ) 2 ) i f t < T m i n T m a x = L t σ t T m i n = σ t σ t = L t 10
s ( t ) = cos 2 ( ω 2 t ) × exp [ ( ω ( t t 0 ) 4 π ) 4 ] ,
Δ v z ( x , z , t ) = Δ φ o p t ( x , z , t ) λ ¯ 4 π n ¯ f s 1 ,
G f i t = a 1 + e b ( x x 0 ) + c
Δ x ( h ) = 2.05 × h + 0.32 mm ,
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.